structure-property relationships directing transport and ... · 1 structure-property relationships...

56
1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand, 1 Sujin Baek, 2 Tzung-Han Lai, 2 Nabankur Deb, 3 Wojciech Zajaczkowski, 4 Romain Stalder, 5 Klaus Müllen, 4 Wojciech Pisula, 4,6 David G. Bucknall, 3 Franky So, 2† John R. Reynolds 1 * 1. School of Chemistry and Biochemistry, School of Materials Science and Engineering, Center for Organic Photonics and Electronics, Georgia Tech Polymer Network, Georgia Institute of Technology, Atlanta, Georgia 30332-0400, United States. 2. Department of Materials Science and Engineering, University of Florida, Gainesville, Florida 32611-7200, United States. 3. School of Materials Science and Engineering, Center for Organic Photonics and Electronics, Georgia Institute of Technology, Atlanta, Georgia 30332-0400, United States. 4. Max Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany. 5. The George and Josephine Butler Polymer Research Laboratory, Department of Chemistry, University of Florida, Gainesville, Florida 32611-7200, United States.

Upload: others

Post on 31-May-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

1

Structure-Property Relationships Directing

Transport and Charge Separation in Isoindigo

Polymers

Caroline Grand,1 Sujin Baek,2 Tzung-Han Lai,2 Nabankur Deb,3 Wojciech Zajaczkowski,4

Romain Stalder,5 Klaus Müllen,4 Wojciech Pisula,4,6 David G. Bucknall,3 Franky So,2† John R.

Reynolds1*

1. School of Chemistry and Biochemistry, School of Materials Science and Engineering, Center

for Organic Photonics and Electronics, Georgia Tech Polymer Network, Georgia Institute of

Technology, Atlanta, Georgia 30332-0400, United States.

2. Department of Materials Science and Engineering, University of Florida, Gainesville, Florida

32611-7200, United States.

3. School of Materials Science and Engineering, Center for Organic Photonics and Electronics,

Georgia Institute of Technology, Atlanta, Georgia 30332-0400, United States.

4. Max Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany.

5. The George and Josephine Butler Polymer Research Laboratory, Department of Chemistry,

University of Florida, Gainesville, Florida 32611-7200, United States.

Page 2: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

2

6. Department of Molecular Physics, Faculty of Chemistry, Lodz University of Technology,

Zeromskiego 116, 90-924 Lodz, Poland.

KEYWORDS. Polymer Solar Cells, Isoindigo Polymers, Bulk Heterojunctions

ABSTRACT. Since being introduced to the open literature in 2010, the isoindigo heterocycle has

been extensively studied as a novel electron deficient building block for organic electronic

materials in conjugated polymers, discrete length oligomers, and molecular systems, particularly

targeting high charge mobility values and ambipolar transport in organic field effect transistors,

along with high power conversion efficiencies in organic photovoltaic devices. This article

introduces results obtained on copolymers of isoindigo with thiophene and alkylated

terthiophenes to highlight fundamental characteristics in isoindigo-based polymers and the

resulting organic field-effect transistors and photovoltaic devices. By comparing and contrasting

the optoelectronic properties, thin film morphology, organic field-effect transistor (OFET)

mobilities, and organic photovoltaic (OPV) performance to previously reported polymers,

structure-processing-property relationships were uncovered. In particular, isoindigo-containing

polymers with more rigid backbones and higher coherence lengths in thin-films lead to increased

charge mobility in OFET devices. In OPV devices, efficiencies over 6% can be obtained by

balancing high ionization potentials typically dictating the open-circuit voltage, and the charge

transfer state energy and blend morphology impacting short-circuit currents. Furthermore, the

impact of polymer structure on solubility and on phase separation in blends with PC71M is

discussed, with isoindigo-based polymers exhibiting lower solubility possibly leading to more

fiber-like morphologies due to both kinetic and thermodynamic effects, either stemming from

Page 3: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

3

polymer dissolution in the casting solvent or from polymer self-assembly during film formation.

This fiber-like polymer morphology remains unaffected by the presence of processing additives,

such as 1,8-diiodooctane. These structure-property relationships developed for isoindigo-based

polymers can also be discussed in the broader context of diketopyrrolopyrrole (DPP) and

thienoisoindigo (TiI) as electron-deficient moieties that can also be doubly substituted on their

amide functionality.

INTRODUCTION. π-Conjugated polymers have been successfully used in solution processed

organic field-effect transistors (OFETs) and organic photovoltaic (OPV) devices, with FET

measured hole mobilities1, 2 on the order of 10 cm2 V-1 s-1 and electron mobilities3, 4 ranging from

0.1 to 1.0 cm2 V-1 s-1,5 while solar cells have attained AM 1.5 power conversion efficiencies

(PCEs) in excess of 9.2%,6-11 and 10.5% in tandem cells.12 In examining this field of materials,

similar polymer structures do not translate from one application to another, but rather need to be

designed to fulfill application specific requirements. In materials designed for OFETs for

instance, there is a need to combine (i) solution processability, (ii) high charge carrier transport,

and (iii) ambient air stability. On the other hand, materials for OPVs may not require record-high

carrier mobilities to perform well, but they also need (iv) broad optical absorption profiles and

high extinction coefficients, as well as (v) appropriate energy levels for charge separation.

Moreover, the solid state order in OFETs,13-16 and phase separation in OPVs are crucial

components towards high performing devices.17

Donor-acceptor (D-A) polymers have been developed using electron accepting units such as

1,3,2-benzothiadiazole (BTD), thienopyrrolodione (TPD) or diketopyrrolopyrrole (DPP) to yield

Page 4: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

4

tunable optoelectronic properties. Recently, the electron deficient isoindigo (iI) unit has shown

promise as an electron acceptor in both OFETs and OPV materials, thanks to its high yielding

synthetic access, two amide functionalities for dual substitution to ensure polymer solubility,

large absorption coefficients and to the suitable electron affinity of the isoindigo polymers to be

used in fullerene-based OPV devices.18 The iI moiety was first reported in a CIBA patent,19 and

subsequently introduced into the open literature by Reynolds et al.20 Since then iI-based

polymers have demonstrated FET hole mobilities21, 22 up to 3 cm2 V-1 s-1 thanks to side-chain

engineering using longer spacers to direct branching points away from the polymer backbone, as

was also reported for other D-A polymers.23 In OPV devices, efficiencies24-27 of 7.0% were

reached using polymer structures based on thienylene-phenylene linkages by tuning the energy

of the frontier molecular orbitals to achieve a balance between a deep electron affinity for

efficient light absorption, charge transfer and charge separation for high short-circuit current

(Jsc), and a deep ionization potential for a high open-circuit voltage (Voc).28 Isoindigo is an

attractive unit for organic electronics as it possesses two amide groups that can be functionalized,

similar to DPP, but in contrast with BTD derivatives, which do not have similar groups for

substitution, and TPD units, which only have one handle for substitution. The possibility of

functionalizing the iI core with solubilizing alkyl, oligoether or siloxane-terminated chains

provides the opportunity to tune optoelectronic and morphological properties, as is the case for

DPP or thienoisoindigo (TiI) units.23

Based on previous literature on iI-based polymers, six fundamental principles are made evident

in these systems with the goal of understanding the correlation between polymer structure, solid-

state morphology, and device properties. While this paper, and the analysis listed below, is

Page 5: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

5

directed to the study of iI-based polymers specifically, similar concepts can be drawn to other bi-

substituted acceptors like DPP and TiI:

- The electron affinity remains around 3.9 eV, as the electronic density in the LUMO is

located on the isoindigo core, while the ionization potential can be tuned by the donor

moiety.29

- Control of the ionization potential leads to tunable open-circuit voltages in OPV devices,

ranging from 0.6 V to 0.9 V depending on the donor selected,18, 30 correlating well with the

effective band gap, i.e. excitation from the ground state of the donor to the excited state of

the acceptor.

- The packing and orientation of the polymer chains can be influenced both by strong

interchain interactions made possible by the rigidity of the isoindigo core compared to

phenylene-flanked DPP, and by the side chain choice.22, 31

- Isoindigo-based polymers can lead to ambipolar transport,32 and high transport

characteristics (up to 3.6 cm2 V-1 s-1 hole mobility,21 and 0.7 cm2 V-1 s-1 electron mobility in

chlorinated isoindigo polymers33). An increase in mobility values is not necessarily

correlated to a decrease in π-stacking distances, and identical characteristic distances can

lead to different mobilities if the morphology is different.

- Polymer solubility can be tuned through the side-chain density along the polymer

backbone as well as through the nature of the side-chains.34, 35 Furthermore, phase

separation between polymer and fullerene in the solid-state is dictated by polymer

solubility in the processing solvent,36 and phase separation is not as drastically modified by

Page 6: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

6

the use of processing additives when polymers exhibit low solubility (<2 mg/mL in ortho-

dichlorobenzene).

- In bulk heterojunctions with PCBM derivatives, polymer-rich domain sizes on the order

of tens of nanometers surrounded by mixed regions of polymer and fullerene can be

achieved through structural design and processing control.37 When coupled to low

resistance in OPV devices this leads to high short-circuit currents and fill factors.

While these principles have been applied to different structures very rarely are they considered

together. To illustrate these concepts, and in particular highlight the donor-acceptor effects on

frontier energy levels and the impact of side-chains on polymer packing and morphology, we

designed a family of three oligothiophene-isoindigo polymers with varying ratios of electron

deficient iI units along the backbone, and variations in the solubilizing side-chain as shown in

Chart 1. The energy levels of the frontier molecular orbitals of an all-thiophene backbone are

modulated by the content of isoindigo units along the conjugated backbone to shift the

absorption to longer wavelengths going from P[T-iI(HD)] to P[T3(C6)-iI(HD)]. Furthermore,

while keeping the conjugated repeat unit constant and following on the work of Bao and co-

workers,22, 38 we designed P[T3(C6)-iI(SiO)] to investigate the effect of the substitution pattern

on solid state order, transport properties and phase separation.

Page 7: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

7

Chart 1. Structures of P[T-iI(HD)], P[T3(C6)-iI(HD)], and P[T3(C6)-iI(SiO)] (red boxes, with

new results discussed in this article), compared to previously published polymers. The color

scheme highlights structures with the same conjugated repeat unit: P[T-iI] – black, P[T2-iI] –

orange; P[DTS-iI] – brown; P[T3-iI] – blue; P[T6-iI] – green.

In this contribution, three new isoindigo polymers are studied in OFET and OPV devices, and

their behavior is compared and contrasted with previously published isoindigo-based polymers

described in Chart 1. These comparative results are used to elucidate a set of guidelines for

molecular design principles for isoindigo containing polymers in organic electronic applications.

The structures discussed investigate the impact of electron-richness along the P[Tn-iI] backbone

on optoelectronic properties, as the electron-rich co-monomer is extended from P[T-iI] (black),

to P[T2-iI] (orange), P[DTS-iI] (brown), P[T3-iI] (blue) and P[T6-iI] (green). With the repeat

unit kept constant, the side-chains on P[Tn(R)-iI(R)] can also be modified to induce changes in

Page 8: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

8

the polymer packing and optoelectronic properties. This study makes detailed comparison of

optoelectronic structure-property relationships to work published on P[T-iI(OD)] from Zhang et

al.,39 P[T-iI(HD)], P[T3(C8)-iI(HD)], and P[T6(C8)-iI(HD)] from Ma, Wang et al.25, 28, 37, 40, 41

and Ho et al.,26 P[T2-iI(OD)] and P[T2-iI(SiO)] from Mei, Kim et al.,22, 42, P[T-iI(OD)], P[T2-

iI(OD)] and P[T2-iI(4-DTd)] from Lei et al.,21, 43 and P[DTS(EH)-iI(EH)] from Stalder et al.44

Grazing incidence wide angle X-ray scattering (GIWAXS) is used to emphasize the role of

microscopic order and backbone orientation on charge transport in OFETs. Density functional

theory (DFT) calculations, along with electrochemical results and charge-modulated

electroabsorption spectroscopy (CMEAS) measurements, were used to understand the charge

separation process and ultimately the open-circuit voltage in OPV devices. Furthermore, AFM

and neutron reflectivity were used to study the lateral and vertical phase separation in the active

layer blends with fullerene derivatives to explain variation in Jsc and fill factor (FF) in OPV

devices.

RESULTS AND DISCUSSION

Polymer synthesis. Isoindigo-based polymers have typically been synthesized using Suzuki and

Stille polycondensations. The synthesis of P[T-iI(HD)]-4 was previously reported,32 and

P[T3(C6)-iI(HD)]-2 and P[T3(C6)-iI(SiO)] were synthesized by Stille polymerization as

described in Scheme S1. We aimed to increase ordering of P[T3-iI] polymers by using n-hexyl

chains on the terthiophene unit rather than the n-octyl chains used previously40 to investigate the

impact of alkyl side-chain length on polymer packing for OFET applications, along with

examining the effect on phase separation with fullerenes in bulk heterojunction (BHJ) OPV

Page 9: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

9

devices. By moving the branching point of the 2-hexyldecyl side chain further away from the

backbone using siloxane side chains, as demonstrated by Mei et al.,45 the impact of polymer

orientation on its interaction with fullerenes can be determined. The number average molecular

weight for these three polymers was estimated by gel permeation chromatography (GPC) in

tetrahydrofuran (THF) and correlates with a degree of polymerization (DP) of 34, 16, and 25 (i.e.

102, 80, and 125 aromatic rings) for P[T-iI(HD)]-4, P[T3(C6)-iI(HD)]-2 and P[T3(C6)-iI(SiO)],

respectively (see Table 1: entries 6, 14 and 19, and SI). The three polymers also have comparable

indices of polydispersity in THF (2.1, 3.7, and 2.2, respectively). Important for comparison,

these macromolecules are similar in size and have reached a point where saturation of their

optoelectronic properties is ensured. In particular, the use of branched side chains on the

isoindigo unit for these polymers ensures solubility during polymerization, and prevents

precipitation at lower molecular weights during polymerization. Furthermore, all three polymers

structures were synthesized to have thienyl end-groups; however, the presence of controlled end-

cappers was not confirmed experimentally.

Theoretical Electronic and Structural Analysis. Estrada et al.29 have previously demonstrated

good correlation between HOMO and LUMO energy levels calculated by density functional

theory (DFT) at the B3LYP/6-31G(d) level and electrochemically determined ionization

potential and electron affinity. As such, DFT calculations at the B3LYP/6-31G(d) level were

carried out on model structures of P[T-iI] and P[T3-iI] (where side chains have been replaced by

methyl groups) to investigate shifts in the energy of the frontier molecular orbitals as a function

of the electron-rich oligothiophene co-monomers. As expected, the less electron-rich T-iI motif

has a deeper calculated HOMO level at -4.72 eV compared to -4.97 eV for T3-iI. This 0.25 eV

difference in HOMO level energy is expected to be reflected in the open circuit voltage (Voc) of

Page 10: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

10

OPV devices with polymers based on a T-iI repeat unit having an increased Voc compared to T3-

iI polymers. In both T-iI and T3-iI, the HOMO is highly delocalized along the polymer chain,

whereas the LUMO is mostly localized on the isoindigo unit as demonstrated by the T3-iI dimer

in Figure 1. The optimized ground state geometry in these model compounds shows a break in

conjugation in the LUMO of P[T3-iI], with electron density located over a single isoindigo unit

and extending slightly over the neighboring thiophene rings. In P[T-iI] systems, the electron

density in the LUMO is spread continuously over two isoindigo cores bridged by a thiophene

ring as seen in Figure S1, which would modify the intrachain transport of electrons and possibly

their interchain hopping characteristics.46, 47 Similar DFT results for T-iI oligomers have been

obtained by De Angelis et al.,48 where the calculated HOMO and LUMO levels (in vacuum) are

within 30 meV of those found in this work, despite a different conformation used for modeling

of T-iI oligomers.

Figure 1. Isodensity surfaces (0.04 e/bohr3) of the frontier orbitals a T3-iI dimer calculated at the

B3LYP/6-31G(d) level: (a) LUMO and (b) HOMO (adapted from 49).

Electrochemically Determined Energy Levels, and Optical Properties. To verify the trends

observed in the energy levels of the frontier molecular orbitals, electrochemical measurements of

the oxidation and reduction potentials were conducted against a ferrocenium/ferrocene (Fc+/Fc)

standard to estimate the ionization potential and electron affinity of the polymers. Based on

Page 11: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

11

measurements of the formal potential for Fc+/Fc at 0.4 V versus a saturated calomel electrode

(SCE) in acetonitrile,50 and of -4.5 eV being equivalent to 0.0 V versus a normal hydrogen

electrode (NHE)51-53 (with SCE at +0.2 V versus NHE ). The formal potential of Fc+/Fc is taken

to be -5.1 eV versus vacuum, which was highlighted by Cardona et al.54 Differential pulse

voltammetry (DPV) measurements on polymer thin films on platinum electrodes in 0.1 M

TBAPF6 in acetonitrile estimated the electron affinity (EA) of the three polymers to be between

3.9 and 3.94 eV. The values match previous reports of the EA of donor-isoindigo-donor

structures being around 3.7 eV and 3.8 eV,18 and are in agreement with the EA for other iI

copolymers listed in Table 1. The ionization potential (IP) of these structures was estimated from

the onset of oxidation. The IP was calculated to be 5.59 eV, 5.57 eV and 5.43 eV for P[T-

iI(HD)]-4, P[T3(C6)-iI(HD)]-2 and P[T3(C6)-iI(SiO)], respectively (entries 6, 14 and 20 in

Table 1). Previously reported values for the HOMO energy level demonstrate that the measured

IP varies with the technique used, with IP of P[T2-iI(OD)] (entry 9) measured at 5.5 eV using

UV-PES and 6.0 eV using cyclic voltammetry with ferrocene oxidation at -5.1 eV against

vacuum.54 In order to support the IP estimated from DPV measurements, UV-PES was also

conducted on thin-films of P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2. The IP was found to be

around 5.2 eV and 5.3 eV in films of P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2 respectively, again

highlighting a 0.4-0.3 eV difference between the IP estimated by DPV relative to the IP

measured by UV-PES. Furthermore, the electron affinity of these polymers is estimated to be

between 3.8 eV and 4.1 eV as expected based on previous DFT calculations, where the

calculated electron density in the LUMO was primarily localized on the iI unit within model

compounds.29 The IP fluctuated over a broader energy range (5.3 eV to 6.1 eV) based on the

Page 12: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

12

electronic nature of the donor moiety, where iI-based polymers with more electron-rich

thiophene units tend to exhibit lower IP’s.

As observed in Figure 2, the absorption spectra of all three isoindigo polymers show a broad

coverage of the visible region, from 400 to 800 nm with absorption maxima around 650 nm.

There are only minor differences observed between the solution and solid-state spectra, where

often times in conjugated polymers a substantial red-shift is observed in the solid state spectra

compared to the spectra in solution. We attribute this to minimal changes in polymer

conformation upon solidification, as both planarization of the polymer chains and interchain

interactions in the solid state lead to a red-shift of the low energy absorption peak or to the

appearance of a low energy shoulder in the UV-vis spectrum as discussed by Kim and Swager.55

Broader UV-vis spectra with a low energy aggregation shoulder are observed for films of

P[T3(C6)-iI(HD)]-2 and P[T3(C6)-iI(SiO)] compared to solution, suggesting an increase in the

intermolecular interactions between neighboring polymer chains in the solid state compared to

solution state. This low energy shoulder is not observed in the case of P[T-iI(HD)]-4 compared

to solution, suggesting a more amorphous arrangement of the polymer backbones in thin films.

This decrease in thin film order for P[T-iI(HD)]-4 is hypothesized to be linked to the higher

density of branched alkyl chains along the backbone, decreasing the extent of interchain

interactions. This discussion based on UV-vis-NIR results is also supported by 2D-NMR

spectroscopy and X-ray scattering data (vide intra). Making comparisons of the many iI-based

polymers reported, all of the entries in Table 1 have an onset of absorption between 1.50 eV and

1.62 eV, likely due to the similar electronic nature of the thiophene-based donors along the

backbone, also described by Stalder et al.18 Figure 2 also highlights that the solution and solid-

Page 13: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

13

state absorption of these oligothiophene-co-iI polymers exhibit two main absorption bands,

which span the visible region of the spectrum to yield blue-green colored polymers.

Table 1. Size exclusion chromatography (SEC), Redox, and Optical Properties of the Isoindigo

Copolymers

SEC UV-vis-NIR Electrochemistry

Entry Polymer Mn

(kDa) PDI

Emax

d

(eV)

solution

Emax

d

(eV)

film

Egap

opt

(eV)

IPe

(eV)

EAe

(eV)

Egap

(eV) Ref

1 P[T-iI(OD)]-1 20 a 2.0 1.76 1.76 1.55 6.1 4.10 2.00 56

2 P[T-iI(OD)]-2 17a 2.1 1.81 1.79 1.58 5.88

39

3 P[T-iI(HD)]-1 86b 2.2 1.78 1.78 1.6 5.82 3.85 1.83 41

4 P[T-iI(HD)]-2 48b 2.7

1.82

5.72 4.00

28

5 P[T-iI(HD)]-3 60b 2.2

1.55 5.75 3.92 1.83 25

6 P[T-iI(HD)]-4 26 a 2.1 1.79 1.79 1.62 5.59/5.21

f 3.90 1.69 32

7 P[T2-iI(OD)]-1 88 a 2.1 1.78 1.78 1.55 6.0 4.0 2.00 56

8 P[T2-iI(OD)]-2 135 a 5.5 1.77 1.77 1.62 5.38

f

22

9 P[T2-iI(OD)]-3 20b 2.0 1.77 1.77 1.60 6.00/5.54

f 4.00 2.00 21

10 P[T2-iI(4-DTd)] 39b 3.2 1.73 1.72 1.58 5.82/5.33

f 4.04 1.78 21

11 P[T2-iI(SiO)] 138 a 3.3 1.70 1.74 1.61 5.20

f

22

12 P[DTS(EH)-iI(EH)] 36 a 2.8 1.71 1.72 1.54 5.55 3.95 1.60 44

13 P[T3(C6)-iI(HD)]-1 24 a 1.9 1.92 1.88 1.58 5.39 3.89 1.50 32

14 P[T3(C6)-iI(HD)]-2 18

a 3.7

1.88 1.90 1.58 5.57/5.29f 3.93 1.64

103c

1.6

15 P[T3(C8)-iI(HD)]-1 43b 3.1 1.91 1.92 1.50 5.79 3.80 1.99 40

16 P[T3(C8)-iI(HD)]-2 73b 2.9

1.92

5.58 3.95

28

17 P[T3(C8)-iI(HD)]-3 100b 3.3

1.50 5.66 3.94 1.72 25

18 P[T3(C8)-iI(HD)]-4 48b 2.4 1.92 1.92 1.59 5.49 3.90 1.59 26

19 P[T3(C6)-iI(SiO)] 32a 2.2 1.89 1.88 1.55 5.43 3.94 1.46

Page 14: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

14

49c

5.4

20 P[T6(C8)-iI(HD)]-1 67b 2.6

1.52 5.44 3.91 1.53 25

21 P[T6(C8)-iI(HD)]-2 35b 1.8 2.00 1.96 1.57 5.37 3.80 1.57 26

Results from this work are highlighted in red. SEC performed in: aTHF, b1,2,4-trichlorobenzene, coCB, all

versus polystyrene standards, dAbsorption maxima in solution and thin film, eionization potential and

electron affinity estimated from CV or DPV homogenized using EIP(EA) = 5.1 + Eonsetox(red) vs Fc+/Fc ;

fHOMO determined by UV-PES. Values in italic obtained from spectra.

Reflecting on the spectroscopic properties of the full set of iI-based polymers in Table 1, it is

seen that isoindigo P[Tn-iI] copolymers show an aggregation shoulder in thin films for n>3

although broadening of the absorption spectra is seen for all the isoindigo polymers listed. Even

in P[T2-iI(SiO)] and P[T2-iI(4-DTd)] films, where the branching point on the side chain is

moved away from the backbone, no aggregation shoulder is observed in the thin film absorption

spectra. As such, based on the UV-vis spectra alone, it would seem that polymers with a lower

density of branched side chains exhibit stronger interchain interactions, to be correlated to 2D-

NMR and X-ray scattering results.

We recently reported the use of solid-state 2D-NMR spectroscopy to probe the correlation

between branched side chain density and interchain interactions.32 The 2D 13C{1H} FSLG-

HETCOR NMR correlation spectra of powder samples of P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-

1 were compared, effectively allowing us to visualize which protons of the branched side chains

contacted strongly with which carbon of the neighboring polymer backbones. From the 2D solid-

state NMR data alone, the higher branched side chain density for P[T-iI(HD)]-4 correlated with

greater overlap of neighboring isoindigo units in the bulk compared to P[T3(C6)-iI(HD)]-1, for

Page 15: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

15

which the 2D NMR spectra suggested stronger intermixing of the isoindigo and terthiophene

portions of the polymer backbone. Given the steric hindrance arising from side chain

functionalization in conjugated polymers, this is consistent with a lower density of branched side

chain affording greater interchain interaction, as seen from the UV-vis data in Figure 2.

300 400 500 600 700 800 900

Film

No

rma

lize

d a

bso

rba

nce

Wavelength (nm)

P[T-iI(HD)]-4

P[T3(C6)-iI(HD)]-2

P[T3(C6)-iI(SiO)]Solution

Figure 2. Normalized UV-vis spectra of P[T-iI(HD)]-4 (black), P[T3(C6)-iI(HD)]-2 (red), and

P[T3(C6)-iI(SiO)] (blue) in chloroform solution (top) and thin film (bottom).

Pristine Polymer Thin Film GIWAXS Characterization. To investigate polymer packing and

the interchain interactions mentioned above, grazing incidence wide angle X-ray scattering

(GIWAXS) measurements were performed on thin films prepared by spin-coating from ortho-

dichlorobenzene (oDCB) solutions onto silicon wafers with a layer of silicon oxide at the

surface. The samples were measured both as-cast and dried at room temperature (GIWAXS

measurements made approximately 7 days after preparation) and after annealing at 200 °C (see

Figure 3). Indexing of the scattering peaks was related to the polymer packing structure, with

(100) planes correlating to lamellar stacking and (010) planes to π-stacking. The interplane

Page 16: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

16

distances were then calculated using the Bragg condition: 𝑑 = 2𝜋𝑄ℎ𝑘𝑙

⁄ .57 After integration over

a quadrant of the detector, line shape analysis was conducted using Gaussian functions to fit the

scattering peaks – the peak positions are attributed to the (hkl) scattering planes and the full

widths at half maximum (FWHM) give indications on the coherence length (CL) calculated by

Scherrer analysis.

Pristine polymer thin films exhibit varying π-stacking distances ranging from 3.6 Å for

P[T3(C6)-iI(SiO)] to 3.8 Å for P[T-iI(HD)]-4, and P[T3(C6)-iI(HD)]-2 has an ordered structure

with a characteristic distance of 4.1 Å that cannot be attributed to π-stacking (Figure 3a-c). This

corroborates what is observed from thin film absorbance measurements with P[T3(C6)-iI(SiO)]

showing the largest red shift among the three polymers. The lack of clear π-stacking scattering

and the larger characteristic distance for P[T3(C6)-iI(HD)]-2 can be explained by increased

rotational freedom in the terthiophene unit containing alkyl chains, compared to a single

thiophene unit in P[T-iI(HD)]-4. The lamellar distance determined from the position of the (100)

peak is similar in films of P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2, around 18 Å, but increases to

25 Å in films of P[T3(C6)-iI(SiO)], possibly due to the longer Si-O-Si distance compared to C-

C-C distance (3.1 Å vs. 2.5 Å, respectively) leading to an overall side chain maximum length of

around 11 Å and 13 Å in the case of 2-hexyldecyl and siloxane side chains respectively. These

values for π-stacking and lamellar distances are close to previous results obtained on

oligothiophene-isoindigo polymers.25 Prior to thermal annealing, the polymer films exhibit

different backbone orientations relative to the substrate. P[T-iI(HD)]-4 and P[T3(C6)-iI(SiO)]

both exhibit (100) lamellar peaks along the Qxy axis (in-plane) and consequently a (010) π-

stacking peak along the Qz axis (polymer backbone mainly face-on relative to the substrate). By

comparison, P[T3(C6)-iI(HD)]-2 has its (100) peak along the Qz axis, and a scattering peak at a

Page 17: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

17

30° angle from the Qxy axis indicating the presence of scattering layers predominantly tilted

towards, and not edge-on to, the substrate.

Figure 3. GIWAXS patterns of pristine (top) and thermally annealed (bottom) P[T-iI(HD)]-4

(a,d), P[T3(C6)-iI(HD)]-2 (b,e) and P[T3(C6)-iI(SiO)] (c,f) thin films prepared from oDCB.

Figures a-e use the same 5,000 intensity scale bar; for optimal data visualization. Figure f uses

the 10,000 intensity scale bar to visualize lower intensity peaks before annealing, without

saturating the scattering intensity after annealing.

The difference in orientation is made even more visible after thermal annealing at 200°C for 5

minutes as shown in Figure 3d-f. In particular in the case of P[T3(C6)-iI(HD)]-2 films, the

scattering peak at 30° off from the Qxy axis (Figure 3b, arrow) disappears, and two new peaks

Page 18: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

18

appear along the Qxy axis after thermal annealing (Figure 3e, arrow). The outermost peak is

attributed to π-stacking of 3.7 Å indicating that the polymer backbones become primarily edge-

on with regards to the substrate after thermal annealing (Table 2, entry 11). The second peak

along the Qxy axis corresponds to a spacing of 4.2 Å, possibly resulting from a rearrangement of

the scattering of scattering peak at 30° off from the Qxy axis. Furthermore, the π-stacking

distance remains at 3.8 Å and 3.6 Å in P[T-iI(HD)]-4 and P[T3(C6)-iI(SiO)] respectively (Table

2: entries 3 and 13 respectively). Thermal annealing also enables a higher degree of order to be

reached in these systems as shown by the higher order scattering peaks in P[T-iI(HD)]-4 and

P[T3(C6)-iI(HD)]-2, and the increase in scattering intensity in films of P[T3(C6)-iI(SiO)]

(Figures 3f uses a different intensity scale than 3e to prevent saturation of the scattering intensity

in the figure). In order to quantify order in the π-stacking direction, Scherrer analysis was used to

calculate CL based on the (010) peak. Interestingly, after thermal annealing the CL of the three

polymers is comparable (around 9-10 nm), corresponding to approximately 20-30 polymer

backbones involved in π-stacks. These numbers are on par with previously reported CL for (010)

peaks for P[T2-iI(OD)]-2 (entry 5), and slightly lower than the length of P[T2-iI(SiO)] (entry 8),

which includes 38 backbones within a π-stack, and higher than P[T3(C8)-iI(HD)]-3 (entry 12),

which includes 8 backbones within a π-stack.

Page 19: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

19

Table 2. Summary of structural characteristics of thermally annealed thin film via GIWAXS,

and OFET device data for the isoindigo polymer series.

Thin-Film Morphology OFET Devices

Entry Polymer

Lamellar

Distance

(Å)

π-

Stacking

(Å)

Backbone

Orientation Architecture

µh µ

e

Ref

(cm2V

-1s

-1)

1 P[T-iI(OD)]-1 22, weak

Relatively

Amorphous

BG/TC,

OTS treateda

1.5×10-

2

NR 56

2 P[T-iI(HD)]-3 20 3.7 Face-on BG/TCb 1×10

-3 8×10

-3

25

3 P[T-iI(HD)]-4

20 3.8 Face-on BG/BC,

HMDS

treatedc

4×10-2

1×10-1

32

20 3.8

CL: 80Å Face-onf

4 P[T2-iI(OD)]-

1 20

Edge-on

BG/TC,

OTS treateda

0.42 NR 56

5 P[T2-iI(OD)]-

2 20

3.8

CL: 80Å Edge-on

BG/TC,

OTS treatedb

0.33 NR 22

6 P[T2-iI(OD)]-

3 20 3.8 Edge-on

BG/TC,

OTS treatedd

0.66 NR 21

7 P[T2-iI(4-

DTd)] 25 3.6 Edge-on

BG/TC,

OTS treatede

2.98 NR 21

8 P[T2-iI(SiO)] 23

3.6

CL:

138Å

Edge- and

face-on

BG/TC,

OTS treatedb

2.00 NR 22

9 P[DTS(EH)-

iI(EH)] 17 3.7 Face-on

BG/BC,

HMDS

treatedc

1×10-3

1×10-3

32

10 P[T3(C6)-

iI(HD)]-1 20 3.7 Face-on

BG/BC,

HMDS

treatedc

5×10-2

- 32

11 P[T3(C6)-

iI(HD)]-2 18

3.7

CL:

100Å

Edge-onf Results on P[T3(C6)-iI(HD)]-1 used

here

Page 20: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

20

12 P[T3(C8)-

iI(HD)]-3 21

3.7

CL: 29Å Edge-on BG/TC

b 3×10

-2 5×10

-3

25

13

P[T3(C6)-

iI(SiO)] 25

3.6

CL: 92Å Face-on

BG/BC,

HMDS

treatedc

4×10-3 -

14 P[T6(C8)-

iI(HD)] 19 3.7 Edge-on BG/TC

b 7×10

-3 2×10

-4

25

aspin coated from chloroform, annealed at 150°C for 20 min; bspin coated from chloroform,

annealed at 170°C for one hour; cdrop cast from chloroform, annealed at 150°C for one hour;

dspin coated from tetrachloroethane, annealed at 150°C for 30 min; espin coated from

tetrachloroethane, annealed at 175°C for 30 min; fspin coated from o-dichlorobenzene, annealed

at 200°C for 5 min. All source/drain contacts are gold. BG: bottom-gate; TC: top-contact; BC:

bottom-contact; CL: coherence length for (010) peak; NR: not reported.

In developing a perspective on these results, trends can be established for this series of

copolymers based on results detailed in Table 2. A first observation is the impact of processing

conditions on the GIWAXS pattern of some polymers. Indeed films of P[T3(C6)-iI(HD)]

processed by drop-casting from a chloroform solution tend to exhibit a (100) lamellar peak along

the Qxy axis (entry 10), whereas films spin coated from oDCB exhibit a (100) peak along the Qz

axis (entry 11). By comparison, P[T-iI(HD)]-4 exhibits a (100) along the Qxy axis, regardless of

the processing conditions. As such, care needs to be made in comparing results between different

studies where structural (repeat unit, degree of polymerization, dispersity, end group identity,

branching, etc.), primary processing (solvent, temperature, substrate, time), and post-processing

(solvent and thermal annealing, use of additives, etc.) all need to be carefully accounted for.

Considering this, a second observation from these results is the lack of correlation between

Page 21: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

21

molecular weight and packing in the case of P[T-iI(HD)] and P[T2-iI(OD)], where variations in

molecular weight do not influence the π-stacking or lamellar distances (entries 2 and 3, and 4-6).

In addition to molecular weight, polydispersity and the identity of end-capping groups are also

expected to impact the scattering results as discussed by Koldemir et al.58 However, the

challenges associated with characterization of the identity of the end-groups and the lack of

reported data on the synthetic addition of end-cappers prevents a discussion on the effect of end-

groups on polymer ordering in thin films in this article. A third observation is that interchain

packing is not influenced by the number of thiophene rings in the backbone, but rather by the

distance between the backbone and branching point on the solubilizing side-chain, with both

siloxane-containing polymers P[T2-iI(SiO)] and P[T3(C6)-iI(SiO)] exhibiting a π-stacking

distance of 3.6 Å (entries 8 and 13). Another noticeable change between the P[T3(C6)-iI(HD)]-2

and P[T3(C6)-iI(SiO)] thin film structures (Figure 3b and c) is their orientation relative to the

substrate, which was also observed in the case of P[T2-iI] by Bao et al.22, 38 Indeed, P[T3(C6)-

iI(SiO)] chains are mainly face-on ((010) π-stacking out-of-plane peak along Qz axis and (100)

lamellar scattering in-plane peak along the Qxy axis), whereas after annealing the P[T3(C6)-

iI(HD)] backbones are predominantly edge-on with regards to the substrate. It was hypothesized

by Chen et al.31 that highly soluble, planar isoindigo-like polymers lead to a majority of the

chains being face-on ((100) lamellar in-plane scattering), which can explain why the more

soluble, less sterically hindered P[T3(C6)-iI(SiO)] shows predominantly face-on orientation

compared to P[T3(C6)-iI(HD)]-2. Another explanation for the face-on orientation of P[T3(C6)-

iI(SiO)] could be the amphiphilic-like effect of the siloxane and alkyl side-chains. As illustrated

by Kim and Swager,55 it is possible that the siloxane side-chains on either side of the isoindigo

units interact preferentially with the SiO2 surface, while the alkyl side-chains on the terthiophene

Page 22: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

22

units are elongated perpendicular to the substrate, leaving the polymer backbone to be face-on

with regards to the substrate. This initial layer of face-on polymer backbone can further act as a

template for the growth of subsequent polymer layers into a face-on arrangement. The different

orientations are expected to affect the charge transport properties of the pristine polymers.

Charge Transport. Solid state structure of thin conjugated polymer films has been shown to

have a large influence on transport properties as measured in field-effect transistors,16 but

straightforward conclusions based on structure alone are difficult to obtain due to the sensitivity

of measurements to processing conditions such as solvent effects (polymer-solvent interactions

and film solidification rate)59 and to the device structure such as substrate choice, electrode

material and contact resistance, and device architecture (bottom gate-bottom contact for e.g.).

With so many variables, care must be taken in comparing these results with the relevant

statistical analysis but general concepts can be extracted from comparing the structures described

in Table 2. To test the influence of polymer packing on transport properties in these iI-based

polymers, bottom-gate, bottom-contact OFETs were fabricated based on P[T3(C6)-iI(SiO)] and

tested, and the results are included in Table 2, which summarizes results previously reported in

the literature on similar isoindigo-based systems. In the approach used here, the active layer was

drop cast at room temperature from a solution of the polymer (2 mg/mL) in chloroform, followed

by drying at 150°C for one hour to reproduce the fabrication conditions of Stalder et al.32

Annealing of the polymer films at 150 °C and 200 °C leads to similar GIWAXS data so both

processing conditions lead to similar observations of polymer packing and conclusions as to the

relationship between film morphology and transport properties.

Previous OFET devices fabricated using these structures have shown ambipolar behavior for

P[T-iI(HD)] (entries 2 and 3), and some conflicting results for P(T3-iI)-like polymers (entries 10

Page 23: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

23

and 12). In the case of P[T3(C6)-iI(HD)]-1, an OFET hole mobility of 5×10-2 cm2V-1s-1 was

obtained, but no n-type behavior was observed. However, when the n-hexyl chains are replaced

by n-octyl chains in P[T3(C8)-iI(HD)]-3, the electron mobility is reported to be 5×10-3 cm2V-1s-1,

with a hole mobility similar to that of P[T3(C6)-iI(HD)]-1 as seen in entries 10 and 12 of Table

2.25 Ambipolar transport was measured in some isoindigo polymers (Table 2, entries 2,3,9,12 and

14); however there is no clear correlation between electron mobility and isoindigo density in the

backbone. This is due in part to the numerous parameters which come into play when measuring

charge mobility in OFET devices. One example where device parameters are kept constant to

focus on the effects of polymer structure on transport was conducted by Stalder et al. using 2D

solid-state NMR to elucidate the interactions between polymer backbones. The authors conclude

that P[T-iI(HD)]-4 packing include slightly shifted isoindigo stacks, which maintain overlap of

the isoindigo units over multiple polymer backbones, whereas P[T3(C6)-iI(HD)]-1 packing

favors mixed interactions of isoindigo and terthiophene across polymer backbones.32 These

differences in molecular packing and orbital overlap between P[T-iI] and P[T3-iI] polymers

account in part for the ambipolar charge transport in P[T-iI] polymer, whereas P[T3-iI] tend to be

p-type polymers. This conclusion is backed up by the fact that the homopolymer of isoindigo,

prepared via Suzuki coupling, exhibits only n-type transport (η = 1×10-4 cm2V-1s-1) due the

propensity of iI interactions.32 In the case of P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-1, the

differences in charge transport can thus be explained by considering intrachain transport at the

molecular level, with breaks in conjugation observed for the LUMO of P[T3-iI], and by

interchain charge hopping through electronic coupling of adjacent iI units, as was discussed in

the case of poly(fluorene-alt-benzothiadiazole) F8BT.46, 47

Page 24: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

24

Examining the results of Table 2, the isoindigo-based polymer films which exhibit π-stacking

have π-π distances of 3.6 Å to 3.8 Å. While a short π-stacking distance may be a necessary

condition for high interchain overlap and high mobilities, P[T-iI(HD)], P[DTS(EH)-iI(HD)], and

P[T3(C6)-iI(SiO)] (entries 2-3, 9 and 13 respectively) demonstrate that it is not a sufficient

characteristic in these polymers. Indeed, all three polymers have a π-stacking distance around 3.7

Å but two also have a face-on polymer orientation relative to the substrate, which sets the charge

hopping direction perpendicular to the transport direction in OFETs, and explains in part their

lower hole mobility compared to P[T3(C6)-iI(HD)]-1. Although orientation is not the only

determining factor in some structures, the increase in mobility with edge-on orientation in

isoindigo polymers seems to follow previous trends observed for DPP-containing polymers.16

Coherence length (as discussed in the previous paragraph, also referred to as crystallite size) in

the π-stacking direction also needs to be considered to explain the trends seen in the transport

properties. For instance, P[T2-iI(OD)] (entries 4-6) shows similar packing to P[T3(C8)-iI(HD)]-3

(entry 12), but possesses a hole mobility that is an order of magnitude higher. This is explained

by the fact that crystallinity is maintained over 3 nm (9 backbone segments) in the case of

P[T3(C8)-iI(HD)]-3 and 8 nm (22 backbone segments) in the case of P[T2-iI(OD)]-2. The

relationship between long-range order in crystallites and high mobility can be used to explain

trends seen in isoindigo polymers, as discussed broadly for conjugated polymers in general in the

recent work by Noriega et al.15 and Zhang et al..14 In one example, Zhang et al. found that both

ordered aggregates and amorphous regions of a copolymer of indacenodithiophene and

benzothiadiazole (IDT-BT) are uniformly oriented with respect to the substrate and both domains

provide pathways for efficient intra- and interchain transport.14 In another publication, Noriega et

al. showed that short-range order within polymer aggregates are connected through amorphous

Page 25: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

25

regions by tie chains, which still allow for efficient interchain transport in the amorphous

domains.15 As such, the work reported here on isoindigo-containing polymers does give some

guidelines for the relationship between coherence length and charge carrier mobility, but our

description is limited in the sense that only the short-range ordered regions were observed by

GIWAXS and no data was obtained to include the impact of amorphous regions of the films on

charge transport.

Solar Cell Properties. To determine the impact of the ground and excited state energy levels,

along with the polymer’s solid state behavior, on photovoltaic properties, solar cells were

fabricated and the processing conditions were optimized using a conventional device structure

(indium tin oxide (ITO)/PEDOT:PSS/Isoindigo-based polymer:PC71BM/LiF/Al).

Page 26: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

26

Table 3. Summary of optimized conventional OPV device characteristics.

Processing OPV Characteristics

Entry Polymer Poly:

PCxBM

Solvent/

Additive

Thick.

(nm)

Jsc

(mA cm-2)

Voc

(V)

FF

(%)

PCE

(%) Ref

1 P[T-iI(OD)]-2 1:2

PC71BM CB 70 1.7 0.87 60 0.9a 39

2

P[T-iI(HD)]-1

1:2

PC71BM

oDCB/

2.5%

DIO

85 5.4 0.89 63 3.0a 41

3 1:1

PC61BM

oDCB/

2.5%

DIO

80 9.1 0.91 54 4.5a 37

4 P[T-iI(HD)]-2 1:1.5

PC61BM

oDCB/

2.5%

DIO

90 5.7 0.92 63 3.3 28

5 P[T-iI(HD)]-3 1:1.5

PC61BM

oDCB/

2.5%

DIO

100 7.2 0.91 66 4.3 a 25

6 P[T-iI(HD)]-4 1:1.5

PC71BM

oDCB/

2.5%

DIO

70 5.8 0.90 55 2.9

7 P[T2-iI(OD)]-2 1:1

PC71BM

CF/

4% DIO 170 10.3 0.91 60 5.2 42

8 P[T2-iI(SiO)] 1:2

PC71BM

CF/

0.5%

DIO

150 9 0.87 52 3.4 42

9 P[DTS(EH)-iI(EH)] 1:4

PC71BM

CB/

4% DIO 8.3 0.76 42 2.6 44

10 P[T3(C6)-iI(HD)]-2 1:1.5

PC71BM

oDCB/

2.5%

DIO

100 15.2 0.70 62 6.6

11 P[T3(C8)-iI(HD)]-1 1:1.5

PC71BM

oDCB/

2.5%

DIO

90 13.1 0.70 69 6.3a 40

12 P[T3(C8)-iI(HD)]-2 1:1.5

PC71BM

oDCB/

2.5%

DIO

90 12.9 0.70 63 5.7 28

Page 27: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

27

13 P[T3(C8)-iI(HD)]-3 1:1.5

PC71BM

oDCB/

2.5%

DIO

100 14.6 0.72 66 6.9 25

14 P[T3(C8)-iI(HD)]-4 1:1.5

PC71BM

oDCB/

3% DIO 13.8 0.73 64 6.4 26

15 P[T3(C6)-iI(SiO)] 1:1.5

PC71BM

oDCB/

2.5%

DIO

90 8.6 0.66 49 2.8

16 P[T6(C8)-iI(HD)]-1 1:1.5

PC71BM

oDCB/

2.5%

DIO

100 10.5 0.65 66 4.7 25

17 P[T6(C8)-iI(HD)]-2 1:1.5

PC71BM

oDCB/

3% CN 15.7 0.71 64 7.1 26

a Best device characteristics, all others are averages.

The optimized device results are summarized in Table 3. As can be seen, the use of 1,8-

dioodoctane (DIO) as a processing additive proves to be beneficial to the PCE in these blends. In

particular the drying time of the active layer was found to have a significant influence on device

performance. When the P[T3(C6)-iI(HD)]-2:PC71BM film was placed under vacuum

immediately after spin-coating the active layer, followed by LiF/Al deposition, the devices

showed a Voc of 0.70±0.01 V, a Jsc of 14.2±0.4 mA cm-2, and FF of 59 %. The PCE was therefore

calculated to be 6.0±0.1 %. However, after drying the active layer overnight at room temperature

in nitrogen atmosphere, the Jsc, FF, and PCE increased to 15.2±0.1 mA cm-2, 62±1.0 %, and

6.6±0.1 %, respectively. These results indicate that an increase in Jsc and FF is observed with the

long evaporation time of the oDCB:DIO mixture. If the evaporation time of the solvent is short,

there is insufficient time for self-organization of the polymer in the film as it solidifies. On the

other hand, long evaporation times allow sufficient time for self-organization to occur.60, 61

Indeed, Rogers et al.62 demonstrated little change in the assembly of poly(2,1,3-

benzothiadiazole-4,7-diyl[4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b:3,4-b′]-dithiophene-2,6-

Page 28: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

28

diyl]) (PCPDTBT) polymer chains in films cast from chlorobenzene past 2 min after casting,

whereas films cast with high boiling point 1,8-octanedithiol (ODT) as a processing additive

showed structural evolution up to 78 min after casting. Furthermore, Ho et al.26 observed

increased short-circuit currents and fill factors in isoindigo-based devices, where the active layer

is processed with 1-chloronaphthalene (CN) as a processing additive. In their work, the active

layer was left to dry overnight under inert atmosphere, which could indicate further

rearrangement with longer drying times when CN is used as a processing additive. One question

that arises is the possibility of the high boiling point additive remaining in the film after casting.

The use of high vacuum (10-6 torr) on the active layer prior to evaporation of the top electrode

provides one mechanism for removal of the processing additive from the film. Previous reports

have discussed the absence of thiol groups from ODT when observed in FTIR, Raman and XPS

after drying the films under 10-3 torr.63 In this work, blends cast with DIO were analyzed by

XPS, and showed no residual iodine peak in the top nanometer layer of the bulk heterojunction

as seen in Table 5. However, complete removal of DIO from the active layer in the final device

is still not fully proven.

As observed in comparing previous reported results summarized in Table 3 to the new results

reported here, we show that the number of thiophenes in the backbone has an influence on the

device performance.25, 26 In the case of P[T-iI(HD)]-4 (entry 6) which has one backbone

thiophene, Voc is 0.9 V while the Voc of P[T3(C6)-iI(HD)]-2 (entry 10) with three backbone

thiophenes is 0.7 V. However, P[T3(C6)-iI(HD)]-2 shows higher PCEs compared to P[T-

iI(HD)]-4 due to higher Jsc and FF. The morphology investigated by atomic force microscopy as

shown in Figure 4 gives some insight as to the cause for this increase in Jsc. As seen in Figures

4b and d, P[T3(C6)-iI(HD)]-2:PC71BM exhibits finer phase separated microstructure compared

Page 29: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

29

to P[T-iI(HD)]-4:PC71BM blends, which is one possible explanation to the higher Jsc in

P[T3(C6)-iI(HD)]-2 devices than P[T-iI(HD)]-4 devices.

Preliminary analysis of Table 3 highlights some discrepancies in Jsc trends between P[T3-iI]

device and P[T6-iI] device as shown by entries 13-14 and 16-17. In work by Ho et al.,26 CN was

used as a solvent additive instead of 1,8-diiodooctane (DIO) that was used by Ma et al.25 in

P[T6(C8)-iI(HD)]:PC71BM blend, leading to an improvement in Jsc from 10.5 mA cm-2 to 15.7

mA cm-2. As summarized in Table 3, almost all reported devices based on isoindigo-containing

polymers required the use of processing additives to achieve optimized OPV performance. As

such, the effect of processing additives on the morphology and device characteristics of P[T-

iI(HD)]-4, P[T3(C6)-iI(HD)]-2 and P[T3(C6)-iI(SiO)]-based films will be discussed in the

following paragraph.

Effect of DIO as a Processing Additive. DIO has been successfully used to control phase

separation in BHJ devices constructed using isoindigo-based polymers blended with fullerenes,

inducing a fine structure, which enhances cell performance.40, 44 As illustrated by the results in

Table 4 for the polymers studied in this work, adding DIO into isoindigo polymer:PC71BM

blends yields an enhancement in Jsc and FF, which in turn leads to higher PCE for all three

polymer systems studied. The largest increase in Jsc with DIO is observed in blends based on

P[T3(C6)-iI(SiO)], with the use of DIO leading to an over five-fold enhancement in Jsc to 8.6 mA

cm-2 compared to 1.6 mA cm-2 without DIO. The Jsc is also increased from 2.3 mA cm-2 to 5.8

mA cm-2 and from 12.5 mA cm-2 to 15.2 mA cm-2 with DIO in blends based on P[T-iI(HD)]-4

and P[T3(C6)-iI(HD)]-2 respectively. Interestingly, the Voc decreases (< 10%) for all devices

cast with DIO compared to devices fabricated without DIO. This was also observed in blends

Page 30: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

30

containing P[DTS(EH)-iI(HD)], where the Voc dropped to 0.76 V with DIO from 0.86 V without

DIO.44

Table 4. Effect of DIO concentration in oDCB on conventional OPV device characteristics

Polymer:PC71BM Jsc

(mA cm-2)

Voc

(V)

FF

(%)

PCE

(%)

P[T-iI(HD)]-4 2.3±0.1 0.97±0.01 52±2 1.2±0.1

P[T-iI(HD)]-4, 2.5% DIO 5.8±0.1 0.90±0.01 55±1 2.9±0.1

P[T3(C6)-iI(HD)]-2 12.5±0.5 0.72±0.01 56±1 5.0±0.2

P[T3(C6)-iI(HD)]-2, 2.5% DIO 15.2±0.1 0.70±0.01 62±1 6.6 ±0.1

P[T3(C6)-iI(SiO)] 1.6±0.3 0.69±0.01 37±1 0.4 ±0.1

P[T3(C6)-iI(SiO)], 2.5% DIO 8.6±0.2 0.66±0.01 49±1 2.8±0.1

Although both P[T3(C6)-iI(HD)]-2 and P[T3(C6)-iI(SiO)] have the same conjugated backbone

structure, they express very different OPV device efficiencies.38 These differences in efficiencies

can be partly explained by the variations in lateral phase structure that cause differences in

charge generation and transport, leading to differences in Jsc and FF. Atomic force microscopy

(AFM) was used to investigate the change in surface morphology as a function of DIO content as

shown in Figure 4. It can be seen that DIO has a large effect on the topology in the case of P[T-

iI(HD)]-4, but little influence on blends with P[T3(C6)-iI(HD)]-2. The finer feature size in the

P[T-iI(HD)]-4 and P[T3(C6)-iI(SiO)] films formed with DIO are consistent with an increase in

Jsc compared to films processed without this additive. Comparing these results, it appears that

DIO as a processing additive has little visible effect on the microscopic phase separation in the

case of materials with low solubility in the processing solvent, such as P[T3(C6)-iI(HD)]-2,

when compared to the more soluble polymers P[T-iI(HD)]-4, P[T3(C6)-iI(SiO)], or P[DTS(EH)-

iI(HD)].44 However, the use of DIO does impact surface roughness in all three polymer films. To

Page 31: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

31

a certain extent, a rougher surface of the active layer in OPV can improve light harvesting and

charge collection. Surface roughness can also be used as an indicator of crystallite formation in

thin films. The increase in surface roughness can explain in part the increase in FF and Jsc, both

leading to higher power conversion efficiency in the devices processed with DIO, although the

correlation between increased surface thickness and increased FF and Jsc remains unclear for

these three systems.

Figure 4. AFM images of blends of PC71BM with (a,b) P[T-iI(HD)]-4, (c,d) P[T3(C6)-iI(HD)]-

2, and (e,f) P[T3(C6)-iI(SiO)] without (top) and with 2.5%v DIO (bottom). The root-mean-

square roughness is indicated at the bottom right corner. All images are 1×1 µm2 with a 30 nm

height scale.

The mechanism by which processing additives modify the blend morphology is complex;64, 65

however, some preliminary work has shown that DIO has an impact on the size of PC71BM

aggregates.66 Without DIO present, fullerene derivatives tend to aggregate into large fullerene

rich domains as the film dries. In comparison, DIO serves as a better solvent for the fullerene

than chlorobenzene or oDCB, and reduces PC71BM aggregation, allowing fullerene to mix more

Page 32: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

32

homogeneously with the polymer in the amorphous phases. In contrast to this hypothesis for

DIO, Schmidt et al. 67 have postulated that CN helps form polymer lamellar crystallites within

polymer aggregates in solution, which in turn leads to smaller and more ordered polymer

domains in the dried thin film. Gao et al.68 also demonstrated the variations in film forming

mechanisms when aliphatic additives (e.g. DIO) are used as opposed to aromatic additives (e.g.

oDCB). In the latter case, polymer aggregates with some degree of ordering are formed in

solution as observed with UV-visible absorption. Yan et al.69 have also discussed the impact of

polymer self-assembly on the choice of processing additive for optimal device performance. CN

was shown to increase Jsc compared to DIO in blends where the polymer exhibited a propensity

to form polymers fibrils, and addition of DIO leads to an increase in Jsc in blends where the

polymer forms less crystalline, irregular particles. This variation in film formation mechanism

could explain the difference in OPV device parameters in the case of P[T6(C8)-iI(HD)] devices

as measured by Ma et al.25 and Ho et al,26 where Jsc of 10.5 mA cm-2 and 15.7 mA cm-2 are

measured in blends cast with DIO and CN respectively (entries 16 and 17). In this case,

differences in self-assembly of P[T3(C8)-iI(HD)] and P[T6(C8)-iI(HD)] chains could explain the

difference in choice of processing additive between DIO and CN for optimized device

performance.

In addition to AFM, GIWAXS was conducted on blends of P[T3(C6)-iI(HD)]-2 and P[T3(C6)-

iI(SiO)] with PC71BM to check for a correlation between molecular ordering and device

parameters (Figures S4 and S5), as discussed in the previous paragraph for other polymer

systems. Based on the GIWAXS results on both polymer:PC71BM blends, slight changes are

observed in the molecular orientation and crystalline packing of the polymers in films spun cast

with and without DIO. In Figure S4, an increase in the scattering intensity is seen in both

Page 33: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

33

polymer blends when DIO is used, suggesting an increase in crystallinity. This is also supported

by the increase in intensity of the higher order (200) peak in blends based on P[T3(C6)-iI(SiO)]

with DIO (Figures S4d and S5c). Another differences observed is a slight broadening of the

(100) peak around the Qz axis when DIO is used (Figure S5, b and d) for both polymers, As

such, the GIWAXS results on polymer crystallites do not indicate a strong dependence of the

molecular morphology of P[T3(C6)-iI(HD)]-2 or P[T3(C6)-iI(SiO)] on the presence of DIO

during processing, but small variations are seen, which can offer a partial explanation for the

increase in FF and Jsc through increased crystallinity when DIO is used.

Beyond lateral phase separation, processing additives can also affect vertical phase separation as

reported by Shao et al.70 in blends based on a copolymer of benzo[1,2-b:4,5-b’]dithiophene and

thieno[3,4-b]thiophene (PBDTTT-C-T). Depending on the vertical phase separation in the

blends, charge collection can be improved at those electrodes by minimizing leakage currents.71

However, in the case of P3HT:PCBM blends, it has been demonstrated that even an interface

made up of only 3% PCBM at the cathode enables low-resistance electron collection.72 In order

to determine the surface composition of the active layer at the PEDOT:PSS and LiF/Al

interfaces, XPS was used on the device active layer to determine the carbon, nitrogen, oxygen

and sulfur content in the top few nanometers of the bulk heterojunction. The relative atomic

content were then related back to the material concentration at the surface using the chemical

formula of the two components in order to quantify the ratio of polymer to fullerene at the

LiF/Al interface of conventional cells,73 as detailed in the supporting information. As seen in

Table 5, there is an increase of the fraction of PC71BM at the surface, near the LiF/Al cathode,

with added DIO.

Page 34: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

34

Table 5. XPS data showing elemental composition and calculated material concentrations at the

film surface as a function of processing conditions.

P[T3(C6)-iI(HD)]-2:PC71BM C

(%)

N

(%)

O

(%)

S

(%)

Polymer

(%)

PC71BM

(%)

No DIO 92.0 2.3 2.8 2.8 83±12 16±10

2.5% DIO 92.5 1.5 3.0 3.0 69±13 30±10

The vertical phase composition was subsequently fully probed using neutron reflectivity (NR) in

order to observe a possible impact of DIO on the vertical composition profile of the active layer.

Figure 5 shows the fits to the experimental data that are used to extract the vertical scattering

length density (SLD) profile for the P[T3(C6)-iI(HD)] blends cast with and without DIO. The

SLD profiles were analyzed in terms of the vertical composition profile of P[T3(C6)-iI(HD)] and

PC71BM, which have calculated SLD values of 7.29×10-7 Å-2 and 4.3×10-6 Å-2 respectively. A

homogeneous blend of these two materials with a 1:1.5 (or 40:60) ratio has a SLD of 2.87×10 -6

Å-2 for the active layer, so the NR fits were also tested to meet this requirement to ensure

accuracy of the fit parameters. Furthermore, the NR fits shown in Figure 5b and c estimate that

the thickness of the bulk heterojunction active layer beneath the aluminum (Al) electrode is

around 125 nm and 100 nm for the blends processed without and with DIO respectively.

Profilometry on the active layer of OPV devices gave active layer thicknesses of 130±3 nm and

120±3 nm for the films processed without and with DIO respectively, which correlates well with

the values obtained from fits to the NR measurements.

In terms of vertical composition in the active layer, (Figure 5b and c) the NR fits indicate that the

active layer has a non-uniform distribution of the polymer and fullerene through the film

Page 35: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

35

thickness. Without added DIO, the surface of the active layer is slightly PC71BM-rich, with a

composition of 52% of PC71BM, and the buried interface at the PEDOT:PSS is polymer-rich,

with a composition of 74% P[T3(C6)-iI(HD)]-2 (decreased SLD) (Table S3). When DIO is used

in the casting solution, the polymer-rich phase at the PEDOT:PSS interface becomes even more

polymer-rich (composition of the layer is 91% P[T3(C6)-iI(HD)]-2) but also becomes thinner (20

nm to a couple nanometers without and with DIO respectively). Based on these NR fits, it is

shown that DIO not only has an impact on lateral phase separation, but also on vertical

composition. However, these NR fits exhibit a different trend from XPS. XPS showed an

increase of the concentration of PC71BM at the LiF/Al interface with DIO, whereas NR fits

indicate a PC71BM-rich interface without DIO. An explanation for this discrepancy is that

capping of the active layer with the cathode material could have an effect on the composition at

that interface. Moreover, the difference in the size of the area analyzed by each technique (4 cm2

in the case of NR, and less than 0.5 mm2 in the case of XPS) could account for variations

between NR and XPS results. Indeed, as seen in Figure S3 and Table S2 of the supporting

information, there is a variation of the surface composition as the measurement is taken closer to

the electrode, where the fraction of polymer drops from 83 % to 70 % closer to the Al layer.

Figure 5. Neutron reflectivity data and fits (a) leading to scattering length density (SLD) profiles

of P[T3(C6)-iI(HD)]-2 devices spun cast without (b) and with 2.5%v DIO (c).

0.01 0.1

1E-7

1E-6

1E-5

1E-4

1E-3

0.01

0.1

1

10

R(l

og

(I))

Q (A-1)

P[T3(C6)-iI(HD)]:PC71BM, no DIO

Fit, no DIO

P[T3(C6)-iI(HD)]:PC71BM, DIO

Fit, DIO

(a)

0 500 1000 1500 2000 2500 3000

5.0x10-7

1.0x10-6

1.5x10-6

2.0x10-6

2.5x10-6

3.0x10-6

3.5x10-6

Active layer: 126 nm

Average SLD: 2.6 x 10-6 A

-2

(b)

Polymer/PC71

BM

40/60

50/50

SiOx

PEDOT:PSS

Polymer rich

Al

SL

D (

A-2)

Distance (A)

P[T3(C6)-iI(HD)]:PC71BM, no DIO

0 500 1000 1500 2000 2500 3000

5.0x10-7

1.0x10-6

1.5x10-6

2.0x10-6

2.5x10-6

3.0x10-6

3.5x10-6

Active layer: 112 nm

Average SLD: 2.7 x 10-6 A

-2

Polymer/PC71

BM

40/60

Polymer rich

SiOx

SL

D (

A-2)

Distance (A)

P[T3(C6)-iI(HD)]:PC71BM, DIO

Al

PEDOT:PSS

50/50

(c)

Page 36: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

36

Balance between Voc and Jsc. Interestingly, devices based on P[T-iI(HD)]-4 exhibit high open-

circuit voltages, but low short-circuit currents, while the opposite is true for P[T3(C6)-iI(HD)]-2.

In general, as described in Table 3, high Voc values of up to 0.91 V are obtained with blends

containing less electron-rich polymer donors (Table 3, entries 1-8), where the high Voc without

sacrificing Jsc leads to 4.5% power conversion efficiency (Table 3, entry 3).25 On the other hand,

in a device containing more electron-rich isoindigo polymer (Table 3, entries 9-17) the Voc was

lowered to 0.7 V while Jsc reached 15.7 mA cm-2, yielding higher PCEs of 7.1 % when CN is

used as a processing additive (Table 3, entry 17).26

As described in previous work,28 device results show a substantial 0.2 V difference in Voc

between P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2 or P[T3(C8)-iI(HD)] devices. Because of the

unstable electrochemical oxidation of the polymers, the understanding of this difference in Voc

based solely on the ionization potential of the polymers is challenging. Moreover, the UV-PES

results show a 0.1 eV difference in the IP of P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2, which does

not explain the 0.25 V difference in the Voc of blends cast without DIO. Vandewal et al.28 have

discussed this difference in the Voc by measuring the energy difference between the excited state

of the donor and the charge transfer state (CT) by Fourier-transform photoelectron spectroscopy

and electroluminescence.28 They find that P[T-iI(HD)]-2:PC61BM blends exhibit a lower energy

offset between the polymer excited state and the CT state (close to 0 eV) compared to P[T3(C8)-

iI(HD)]:PC71BM blends, along with low radiative charge transfer state recombination. This leads

to low voltage losses in P[T-iI(HD)]-2 cells, resulting in a high Voc of 0.9V in the case of P[T-

iI(HD)]:PC61BM blends. Furthermore, field-dependent photoluminescence quenching indicates

that this low energy difference between excited and CT states in P[T-iI(HD)]-2:PC61BM blends

Page 37: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

37

is responsible for the low Jsc due to repopulation of the exciton from the CT state, rather than

charge separation.

In this current work, charge modulated electroabsorption spectroscopy (CMEAS) is used to

understand the change in Voc between P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2. CMEAS is a

technique that detects the changes in optical absorption below the polymer bandgap by

modulating the electric field to directly measure the effective band gap in a bulk heterojunction

solar cell.74 This novel technique allows detection of the CT states which have been correlated

with Voc. CMEAS provides higher signal-to-noise resolution compared to linear optical

absorption techniques, and takes into account interfacial effects between the two compounds in

contrast to electrochemical measurements.

The results obtained by Lai et al. using CMEAS show that the effective energy gap in P[T-

iI(HD)]-4:PC71BM blends is measured to be 1.3 eV compared to 1.0 eV in P[T3(C6)-iI(HD)]-

2:PC71BM blends.49 The 0.3 eV difference between these two blends is in agreement with the

variations observed for the Voc of the OPV devices. The CMEAS results indicate that, while

polymers may have similar onsets of oxidation, the CT states formed at the donor/acceptor

interface can have different energy levels. This difference in CT energy can be explained by the

differences in electronic coupling at the interface between polymer and fullerene in blends of

P[T-iI(HD)]-4:PC71BM and P[T3(C6)-iI(HD)]-2:PC71BM respectively, as described for other

systems by Graham et al,75 which affects the effective energy gap. The difference between CT

state energy and Voc is further explained by energy losses to reach charge separated states from

the CT state. Furthermore, higher dissociation efficiency is observed as a result of the more

delocalized CT states in P[T3(C6)-iI(HD)]-2:PC71BM blends than P[T-iI(HD)]-4:PC71BM

blends, resulting in a higher Jsc in solar cells.76 This underlines the role played by dielectric

Page 38: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

38

constants in isoindigo-based blends, which could also explain why the low energy offset between

excited and CT states still leads to high internal quantum efficiencies (IQE) in the case of

P[T3(C6)-iI(HD)] or P[T3(C8)-iI(HD)].

Inverted Device Architecture. The effect of device architecture on device properties was

further investigated in inverted devices, more amenable to roll-to-roll fabrication. Inverted bulk-

heterojunction solar cells with a structure of ITO glass/ZnO–PVP nanocomposite/isoindigo-

based polymer:PC71BM /MoOx/Ag was fabricated using the three focus polymers in this work:

P[T-iI(HD)]-4, P[T3(C6)-iI(HD)]-2, and P[T3(C6)-iI(SiO)]. We previously reported ZnO–

poly(vinyl pyrrolidone) (PVP) composite sol–gel film5 as the ETL. On top of the ZnO–PVP

composites, the active layer in inverted devices was deposited using the same conditions as in

conventional devices. The inverted device results are summarized in Figure 6 and Table 6. For

both P[T-iI(HD)]-4 and P[T3(C6)-iI(HD)]-2, the conventional and inverted devices showed

similar device performances. On the other hand, inverted devices based on P[T3(C6)-iI(SiO)]

showed higher device performance, such as higher Jsc and FF in spite of having a lower Voc,

compared to conventional devices. The increase in device efficiency in P[T3(C6)-

iI(SiO)]:PC71BM inverted OPV devices could not only result from a difference in electrode work

function, but also from variations in the blend morphology and carrier recombination

dynamics.77

Page 39: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

39

Figure 6. J-V curves of conventional and inverted solar cells based on P[T-iI(HD)]-4, P[T3(C6)-

iI(HD)]-2, and P[T3(C6)-iI(SiO)] polymer:PC71BM blends.

Table 6. Conventional and inverted device characteristics.

Device Jsc

(mA cm-2)

Voc

(V)

FF

(%)

PCE

(%)

P[T-iI(HD)]-4 conventional 5.8±0.1 0.90±0.01 55±1 2.9±0.1

P[T-iI(HD)]-4 inverted 5.6±0.1 0.81±0.01 52±1 2.4±0.1

P[T3(C6)-iI(HD)]-2 conventional 15.2±0.1 0.70±0.01 62±1 6.6±0.1

P[T3(C6)-iI(HD)]-2 inverted 15.2±0.3 0.68±0.01 63±1 6.5±0.2

P[T3(C6)-iI(SiO)] conventional 8.6±0.2 0.66±0.01 49±1 2.8±0.1

P[T3(C6)-iI(SiO)] inverted 9.5±0.1 0.64±0.01 63±1 3.8±0.1

SUMMARY AND PERSPECTIVE

-0.2 0.0 0.2 0.4 0.6 0.8 1.0

-16

-12

-8

-4

0

4

8

12

16C

urr

en

t d

en

sit

y (

mA

/cm

2)

Voltage (V)

P[T-iI(HD)], Conventional

P[T-iI(HD)], Inverted

P[T3(C6)-iI(HD)], Conventional

P[T3(C6)-iI(HD)], Inverted

P[T3(C6)-iI(SiO)], Conventional

P[T3(C6)-iI(SiO)], Inverted

Page 40: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

40

In summary, we have synthesized and characterized a series of new isoindigo polymers to

highlight structure-processing-property relationships which can lead others in further developing

this family of polymers. Facile access to isoindigo in bulk syntheses suggests these systems are

worth optimization and bulk preparation for further practical development of OPV’s. These

polymers were designed to include one or three thiophene units as electron-rich moieties, and to

study the impact of side-chain branching on the electronic properties of the resulting polymer

and on the solid-state morphology. In terms of electronic and geometric considerations, the

strong electron accepting character of isoindigo localizes the electron density in the LUMO and

sets the electron affinity around 3.9 eV. The impact of side-chains on these interchain

interactions was discussed using UV-Vis absorption, 2D NMR and GIWAXS data, where

increasing the density of branched side-chains along the backbone tends to decrease interchain

interactions. These interchain interactions are particularly important for reaching high mobility

values in OFETs, where charge transport was shown to be consistently improved by increasing

the order of the isoindigo-containing polymer in thin-films, as quantified here by the coherence

length. This can be achieved by promoting backbone interactions through reducing steric

hindrance by moving the side chain branching point away from the backbone as shown in this

article by the P[T2-iI] family, and in other literature reports.78, 79 Orientation of polymer chains in

the bulk of thin films can be dictated by the chemical nature and density of the side chains, with

alkylated side chains mainly leading to edge-on orientations compared to siloxane side chains,

which yield some face-on orientation. This leads to increased transport in OFETs when the iI-

containing polymer backbones are preferentially oriented perpendicular to the substrate, and the

charge carrier mobility depends on the coherence length rather than on the π-stacking distance in

these isoindigo-based systems. Furthermore, the interchain interactions are crucial to the type of

Page 41: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

41

charge carrier transport and to the magnitude of the charge carrier mobility, where 2D NMR was

used to correlate the iI-iI interchain interactions in P[T-iI] polymers to ambipolar charge

transport, whereas T3-iI interactions were prevalent in P[T3-iI] polymers leading to p-type

charge transport.

In OPV devices, high Voc can be achieved thanks to high ionization potentials of the isoindigo-

based polymers, which needs to be balanced with a high density of charge transfer states to

achieve high Jsc, and thus optimal PCE. The morphology in isoindigo-based blends can be tuned

by using solvent additives such as DIO or CN; however, in the case of P[T3(C6)-iI(HD)]-2 the

use of DIO does not lead to a large change in phase separation as is the case for P[T-iI(HD)]-4 or

P[T3(C6)-iI(SiO)] as investigated by AFM. The increased solubility of P[T-iI(HD)]-4 and

P[T3(C6)-iI(SiO)] in oDCB compared to P[T3(C6)-iI(HD)] is one explanation for the more

visible change in morphology when a poor solvent for the polymer, such as DIO, is used. These

observations point to the impact of the state of the polymer chains in solution prior to film

formation on the resulting thin-film morphology, in addition to the impact of residual of high

boiling point solvents in the film to allow for polymer chain rearrangement or fullerene

aggregation. This work, along with previous reports,25, 26, 40 also highlights the balance in OPV

performance, with PCEs over 6% in both conventional and inverted devices with iI-based

polymers, and the synthetic accessibility of isoindigo-containing polymers.

Overall, this work establishes some structure-processing-property relationships that could lead to

the design of novel high performance isoindigo-based polymers. Furthermore, this work can be

extended to other dually substituted acceptors such as DPP or TiI, where most of the differences

between these three units stem from their electronic characteristics, where iI-based polymers

tend to exhibit the highest ionization potential and electron affinity.

Page 42: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

42

ASSOCIATED CONTENT

Supporting Information. Experimental details, electrochemical details and square-wave

voltamograms, description, OFET and OPV device fabrication, XPS sample preparation and

analysis, additional data from DFT calculation, monomer and polymer synthesis and gel

permeation chromatograms are supplied as Supporting Information. This material is available

free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION

Corresponding Author

* Email: [email protected]

Present Addresses

† Current address: Department of Materials Science and Engineering, North Carolina State

University, Raleigh, NC, USA

Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval

to the final version of the manuscript.

Funding Sources

JRR acknowledges funding of this work from the Office of Naval Research (N00014-14-1-

0173).

Notes

The authors declare no competing financial interest.

Page 43: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

43

ACKNOWLEDGMENT

JRR acknowledges funding of this work from the Office of Naval Research (N00014-14-1-

0173). CG acknowledges the Center for Organic Photonics and Electronics for a fellowship and

R. S. acknowledges the University Alumni Awards Program for a fellowship.

REFERENCES

1. Li, J.; Zhao, Y.; Tan, H. S.; Guo, Y.; Di, C.-A.; Yu, G.; Liu, Y.; Lin, M.; Lim, S. H.;

Zhou, Y.; Su, H.; Ong, B. S. A Stable Solution-Processed Polymer Semiconductor with Record-

High Mobility for Printed Transistors. Sci. Rep. 2012, 2, 754.

2. Kim, G.; Kang, S.-J.; Dutta, G. K.; Han, Y.-K.; Shin, T. J.; Noh, Y.-Y.; Yang, C. A

Thienoisoindigo-Naphthalene Polymer with Ultrahigh Mobility of 14.4 cm2/V·s That

Substantially Exceeds Benchmark Values for Amorphous Silicon Semiconductors. J. Am. Chem.

Soc. 2014, 136, 9477-9483.

3. Lee, J.; Han, A. R.; Yu, H.; Shin, T. J.; Yang, C.; Oh, J. H. Boosting the Ambipolar

Performance of Solution-Processable Polymer Semiconductors via Hybrid Side-Chain

Engineering. J. Am. Chem. Soc. 2013, 135, 9540-9547.

4. Li, H.; Kim, F. S.; Ren, G.; Jenekhe, S. A. High-Mobility n-type Conjugated Polymers

Based on Electron-Deficient Tetraazabenzodifluoranthene Diimide for Organic Electronics. J.

Am. Chem. Soc. 2013, 135, 14920-14923.

5. Holliday, S.; Donaghey, J. E.; McCulloch, I. Advances in Charge Carrier Mobilities of

Semiconducting Polymers Used in Organic Transistors. Chem. Mater. 2014, 26, 647-663.

Page 44: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

44

6. Liang, Y.; Yu, L. A New Class of Semiconducting Polymers for Bulk Heterojunction

Solar Cells with Exceptionally High Performance. Acc. Chem. Res. 2010, 43, 1227-1236.

7. Small, C. E.; Chen, S.; Subbiah, J.; Amb, C. M.; Tsang, S.-W.; Lai, T.-H.; Reynolds, J.

R.; So, F. High-Efficiency Inverted Dithienogermole-Thienopyrrolodione-Based Polymer Solar

Cells. Nature Photon. 2012, 6, 115-120.

8. He, Z.; Zhong, C.; Su, S.; Xu, M.; Wu, H.; Cao, Y. Enhanced Power-Conversion

Efficiency in Polymer Solar Cells Using an Inverted Device Structure. Nature Photon. 2012, 6,

591-595.

9. He, Z.; Xiao, B.; Liu, F.; Wu, H.; Yang, Y.; Xiao, S.; Wang, C.; Russell, T. P.; Cao, Y.

Single-Junction Polymer Solar Cells with High Efficiency and Photovoltage. Nature Photon.

2015, 9, 174-179.

10. Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin, H.; Ade, H.; Yan, H.

Aggregation and Morphology Control Enables Multiple Cases of High-Efficiency Polymer Solar

Cells. Nature Commun. 2014, 5, 5293.

11. Hu, H.; Jiang, K.; Yang, G.; Liu, J.; Li, Z.; Lin, H.; Liu, Y.; Zhao, J.; Zhang, J.; Huang,

F.; Qu, Y.; Ma, W.; Yan, H. Terthiophene-Based D–A Polymer with an Asymmetric

Arrangement of Alkyl Chains That Enables Efficient Polymer Solar Cells. J. Am. Chem. Soc.

2015, 137, 14149-14157.

12. You, J.; Dou, L.; Yoshimura, K.; Kato, T.; Ohya, K.; Moriarty, T.; Emery, K.; Chen, C.-

C.; Gao, J.; Li, G.; Yang, Y. A Polymer Tandem Solar Cell with 10.6% Power Conversion

Efficiency. Nature Commun. 2013, 4, 1446-1446.

Page 45: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

45

13. Mei, J.; Diao, Y.; Appleton, A. L.; Fang, L.; Bao, Z. Integrated Materials Design of

Organic Semiconductors for Field-Effect Transistors. J. Am. Chem. Soc. 2013, 135, 6724-6746.

14. Zhang, X.; Bronstein, H.; Kronemeijer, A. J.; Smith, J.; Kim, Y.; Kline, R. J.; Richter, L.

J.; Anthopoulos, T. D.; Sirringhaus, H.; Song, K.; Heeney, M.; Zhang, W.; McCulloch, I.;

DeLongchamp, D. M. Molecular Origin of High Field-Effect Mobility in an

Indacenodithiophene–Benzothiadiazole Copolymer. Nature Commun. 2013, 4, 2238.

15. Noriega, R.; Rivnay, J.; Vandewal, K.; Koch, F. P. V.; Stingelin, N.; Smith, P.; Toney,

M. F.; Salleo, A. A General Relationship Between Disorder, Aggregation and Charge Transport

in Conjugated Polymers. Nature Mater. 2013, 12, 1038–1044.

16. Sirringhaus, H. 25th Anniversary Article: Organic Field-Effect Transistors: The Path

Beyond Amorphous Silicon. Adv. Mater. 2014, 26, 1319-35.

17. Dang, M. T.; Hirsch, L.; Wantz, G.; Wuest, J. D. Controlling the Morphology and

Performance of Bulk Heterojunctions in Solar Cells. Lessons Learned from the Benchmark

Poly(3-hexylthiophene):[6,6]-phenyl-C61-butyric Acid Methyl Ester System. Chem. Rev. 2013,

113, 3734-65.

18. Stalder, R.; Mei, J.; Graham, K. R.; Estrada, L. A.; Reynolds, J. R. Isoindigo, a Versatile

Electron-Deficient Unit For High-Performance Organic Electronics. Chem. Mater. 2014, 26,

664-678.

19. Flores, J.-C.; Berens, U.; Bienewald, F.; Kirnerand, H. J.; Turbiez, M. G. R. (CIBA)

Ketopyrroles as organic semiconductors. PCT Int. Appl. WO 2009053291 A1, 2009; 150:

495343.

Page 46: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

46

20. Mei, J.; Graham, K. R.; Stalder, R.; Reynolds, J. R. Synthesis of Isoindigo-Based

Oligothiophenes for Molecular Bulk Heterojunction Solar Cells. Org. Lett. 2010, 12, 660-663.

21. Lei, T.; Dou, J.-H.; Pei, J. Influence of Alkyl Chain Branching Positions on the Hole

Mobilities of Polymer Thin-Film Transistors. Adv. Mater. 2012, 24, 6457-6461.

22. Mei, J.; Kim do, H.; Ayzner, A. L.; Toney, M. F.; Bao, Z. Siloxane-Terminated

Solubilizing Side Chains: Bringing Conjugated Polymer Backbones Closer and Boosting Hole

Mobilities in Thin-Film Transistors. J. Am. Chem. Soc. 2011, 133, 20130-3.

23. Mei, J.; Bao, Z. Side Chain Engineering in Solution-Processible Conjugated Polymers for

Organic Solar Cells and Field-Effect Transistors. Chem. Mater. 2014, 26, 604-615.

24. Deng, Y.; Liu, J.; Wang, J.; Liu, L.; Li, W.; Tian, H.; Zhang, X.; Xie, Z.; Geng, Y.;

Wang, F. Dithienocarbazole and Isoindigo Based Amorphous Low Bandgap Conjugated

Polymers for Efficient Polymer Solar Cells. Adv. Mater. 2014, 26, 471-476.

25. Ma, Z.; Sun, W.; Himmelberger, S.; Vandewal, K.; Tang, Z.; Bergqvist, J.; Salleo, A.;

Andreasen, J. W.; Inganas, O.; Andersson, M. R.; Muller, C.; Zhang, F.; Wang, E. Structure-

Property Relationships of Oligothiophene-Isoindigo Polymers for Efficient Bulk-Heterojunction

Solar Cells. Energy Environ. Sci. 2014, 7, 361-369.

26. Ho, C.-C.; Chen, C.-A.; Chang, C.-Y.; Darling, S. B.; Su, W.-F. Isoindigo-Based

Copolymers for Polymer Solar Cells with Efficiency over 7%. J. Mater. Chem. A 2014, 2, 8026-

8032.

Page 47: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

47

27. Dong, X.; Deng, Y.; Tian, H.; Xie, Z.; Geng, Y.; Wang, F. Isoindigo-Based Low

Bandgap Conjugated Polymer for o-Xylene Processed Efficient Polymer Solar Cells with Thick

Active Layers. J. Mater. Chem. A 2015, 3, 19928-19935.

28. Vandewal, K.; Ma, Z.; Bergqvist, J.; Tang, Z.; Wang, E.; Henriksson, P.; Tvingstedt, K.;

Andersson, M. R.; Zhang, F.; Inganäs, O. Quantification of Quantum Efficiency and Energy

Losses in Low Bandgap Polymer:Fullerene Solar Cells with High Open-Circuit Voltage. Adv.

Funct. Mater. 2012, 22, 3480-3490.

29. Estrada, L. A.; Stalder, R.; Abboud, K. A.; Risko, C.; Brédas, J.-L.; Reynolds, J. R.

Understanding the Electronic Structure of Isoindigo in Conjugated Systems: A Combined

Theoretical and Experimental Approach. Macromolecules 2013, 46, 8832-8844.

30. Wang, E.; Mammo, W.; Andersson, M. R. 25th Anniversary Article: Isoindigo-Based

Polymers and Small Molecules for Bulk Heterojunction Solar Cells and Field Effect Transistors.

Adv. Mater. 2014, 26, 1801-1826.

31. Chen, M. S.; Niskala, J. R.; Unruh, D. A.; Chu, C. K.; Lee, O. P.; Fréchet, J. M. J.

Control of Polymer-Packing Orientation in Thin Films through Synthetic Tailoring of Backbone

Coplanarity. Chem. Mater. 2013, 25, 4088-4096.

32. Stalder, R.; Puniredd, S. R.; Hansen, M. R.; Koldemir, U.; Grand, C.; Zajaczkowski, W.;

Müllen, K.; Pisula, W.; Reynolds, J. R. Ambipolar Charge Transport in Isoindigo-Based Donor–

Acceptor Polymers. Chem. Mater. 2016, 28, 1286-1297.

33. Lei, T.; Dou, J.-H.; Ma, Z.-J.; Liu, C.-J.; Wang, J.-Y.; Pei, J. Chlorination as a Useful

Method to Modulate Conjugated Polymers: Balanced and Ambient-Stable Ambipolar High-

Page 48: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

48

Performance Field-Effect Transistors and Inverters Based on Chlorinated Isoindigo Polymers.

Chem. Sci. 2013, 4, 2447-2452.

34. Cao, K.; Wu, Z.; Li, S.; Sun, B.; Zhang, G.; Zhang, Q. A Low Bandgap Polymer Based

on Isoindigo and Bis(dialkylthienyl)benzodithiophene for Organic Photovoltaic Applications. J.

Polym. Sci. A Polym. Chem. 2013, 51, 94-100.

35. Xu, X.; Cai, P.; Lu, Y.; Choon, N. S.; Chen, J.; Hu, X.; Ong, B. S. Synthesis and

Characterization of Thieno[3,2-b]thiophene-Isoindigo-Based Copolymers as Electron Donor and

Hole Transport Materials for Bulk-Heterojunction Polymer Solar Cells. J. Polym. Sci. A Polym.

Chem. 2013, 51, 424-434.

36. Li, W.; Hendriks, K. H.; Furlan, A.; Roelofs, W. S. C.; Wienk, M. M.; Janssen, R. A. J.

Universal Correlation between Fibril Width and Quantum Efficiency in Diketopyrrolopyrrole-

Based Polymer Solar Cells. J. Am. Chem. Soc. 2013, 135, 18942-18948.

37. Ma, Z.; Wang, E.; Vandewal, K.; Andersson, M. R.; Zhang, F. Enhance Performance of

Organic Solar Cells Based on an Isoindigo-Based Copolymer by Balancing Absorption and

Miscibility of Electron Acceptor. Appl. Phys. Lett. 2011, 99, 143302-143304.

38. Kim, D. H.; Ayzner, A. L.; Appleton, A. L.; Schmidt, K.; Mei, J.; Toney, M. F.; Bao, Z.

Comparison of the Photovoltaic Characteristics and Nanostructure of Fullerenes Blended with

Conjugated Polymers with Siloxane-Terminated and Branched Aliphatic Side Chains. Chem.

Mater. 2013, 25, 431-440.

39. Zhang, G.; Fu, Y.; Xie, Z.; Zhang, Q. Synthesis and Photovoltaic Properties of New Low

Bandgap Isoindigo-Based Conjugated Polymers. Macromolecules 2011, 44, 1414-1420.

Page 49: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

49

40. Wang, E.; Ma, Z.; Zhang, Z.; Vandewal, K.; Henriksson, P.; Inganas, O.; Zhang, F.;

Andersson, M. R. An Easily Accessible Isoindigo-Based Polymer for High-Performance

Polymer Solar Cells. J. Am. Chem. Soc. 2011, 133, 14244-14247.

41. Wang, E.; Ma, Z.; Zhang, Z.; Henriksson, P.; Inganas, O.; Zhang, F.; Andersson, M. R.

An Isoindigo-Based Low Band Gap Polymer for Efficient Polymer Solar Cells with High Photo-

Voltage. Chem. Commun. 2011, 47, 4908-4910.

42. Kim, D. H.; Ayzner, A. L.; Appleton, A. L.; Schmidt, K.; Mei, J.; Toney, M. F.; Bao, Z.

Comparison of the Photovoltaic Characteristics and Nanostructure of Fullerenes Blended with

Conjugated Polymers with Siloxane-Terminated and Branched Aliphatic Side Chains. Chem.

Mater. 2012, 25, 431-440.

43. Lei, T.; Cao, Y.; Zhou, X.; Peng, Y.; Bian, J.; Pei, J. Systematic Investigation of

Isoindigo-Based Polymeric Field-Effect Transistors: Design Strategy and Impact of Polymer

Symmetry and Backbone Curvature. Chem. Mater. 2012, 24, 1762-1770.

44. Stalder, R.; Grand, C.; Subbiah, J.; So, F.; Reynolds, J. R. An Isoindigo and Dithieno[3,2-

b:2',3'-d]silole Copolymer for Polymer Solar Cells. Polym. Chem. 2012, 3, 89-92.

45. Ko, S.; Hoke, E. T.; Pandey, L.; Hong, S.; Mondal, R.; Risko, C.; Yi, Y.; Noriega, R.;

McGehee, M. D.; Brédas, J.-L.; Salleo, A.; Bao, Z. Controlled Conjugated Backbone Twisting

for an Increased Open-Circuit Voltage while Having a High Short-Circuit Current in

Poly(hexylthiophene) Derivatives. J. Am. Chem. Soc. 2012, 134, 5222–5232.

46. Donley, C. L.; Zaumseil, J.; Andreasen, J. W.; Nielsen, M. M.; Sirringhaus, H.; Friend,

R. H.; Kim, J.-S. Effects of Packing Structure on the Optoelectronic and Charge Transport

Page 50: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

50

Properties in Poly(9,9-di-n-octylfluorene-alt-benzothiadiazole). J. Am. Chem. Soc. 2005, 127,

12890-12899.

47. Van Vooren, A.; Kim, J.-S.; Cornil, J. Intrachain versus Interchain Electron Transport in

Poly(fluorene-alt-benzothiadiazole): A Quantum-Chemical Insight. ChemPhysChem 2008, 9,

989-993.

48. Salvatori, P.; Mosconi, E.; Wang, E.; Andersson, M.; Muccini, M.; De Angelis, F.

Computational Modeling of Isoindigo-Based Polymers Used in Organic Solar Cells. J. Phys.

Chem. C 2013, 117, 17940-17954.

49. Lai, T.-H.; Constantinou, I.; Grand, C. M.; Klump, E. D.; Baek, S.; Hsu, H.-Y.; Tsang,

S.-W.; Schanze, K. S.; Reynolds, J. R.; So, F. Evidence of Molecular Structure Dependent

Charge Transfer between Isoindigo-Based Polymers and Fullerene. Chem. Mater. 2016, DOI:

10.1021/acs.chemmater.6b00824.

50. Pavlishchuk, V. V.; Addison, A. W. Conversion Constants for Redox Potentials

Measured Versus Different Reference Electrodes in Acetonitrile Solutions at 25°C. Inorg. Chim.

Acta. 2000, 298, 97-102.

51. Trasatti, S. The Absolute Electrode Potential: An Explanatory Note. Pure Appl. Chem.

1986, 58, 955-966.

52. Hansen, W. N.; Hansen, G. J. Absolute Half-Cell Potential: A Simple Direct

Measurement. Phys. Rev. A 1987, 36, 1396-1402.

Page 51: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

51

53. Bard, A. J.; Faulkner, L. R., Electrochemical Methods: Fundamentals and Applications,

2nd ed.; Wiley: New York, 2001.

54. Cardona, C. M.; Li, W.; Kaifer, A. E.; Stockdale, D.; Bazan, G. C. Electrochemical

Considerations for Determining Absolute Frontier Orbital Energy Levels of Conjugated

Polymers for Solar Cell Applications. Adv. Mater. 2011, 23, 2367-2371.

55. Kim, J.; Swager, T. M. Control of Conformational and Interpolymer Effects in

Conjugated Polymers. Nature 2001, 411, 1030-1034.

56. Lei, T.; Cao, Y.; Fan, Y.; Liu, C. J.; Yuan, S. C.; Pei, J. High-Performance Air-Stable

Organic Field-Effect Transistors: Isoindigo-Based Conjugated Polymers. J. Am. Chem. Soc.

2011, 133, 6099-6101.

57. Rivnay, J.; Mannsfeld, S. C. B.; Miller, C. E.; Salleo, A.; Toney, M. F. Quantitative

Determination of Organic Semiconductor Microstructure from the Molecular to Device Scale.

Chem. Rev. 2012, 112, 5488-5519.

58. Koldemir, U.; Puniredd, S. R.; Wagner, M.; Tongay, S.; McCarley, T. D.; Kamenov, G.

D.; Müllen, K.; Pisula, W.; Reynolds, J. R. End Capping Does Matter: Enhanced Order and

Charge Transport in Conjugated Donor–Acceptor Polymers. Macromolecules 2015, 48, 6369-

6377.

59. Ho, P. K. H.; Chua, L. L.; Dipankar, M.; Gao, X. Y.; Qi, D. C.; Wee, A. T. S.; Chang, J.

F.; Friend, R. H. Solvent Effects on Chain Orientation and Interchain π-Interaction in Conjugated

Polymer Thin Films: Direct Measurements of the Air and Substrate Interfaces by Near-Edge X-

ray Absorption Spectroscopy. Adv. Mater. 2007, 19, 215-221.

Page 52: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

52

60. Shin, N.; Richter, L. J.; Herzing, A. A.; Kline, R. J.; DeLongchamp, D. M. Effect of

Processing Additives on the Solidification of Blade-Coated Polymer/Fullerene Blend Films via

In-Situ Structure Measurements. Adv. Energy Mater. 2013, 3, 938-948.

61. Richter, L. J.; DeLongchamp, D. M.; Bokel, F. A.; Engmann, S.; Chou, K. W.; Amassian,

A.; Schaible, E.; Hexemer, A. In Situ Morphology Studies of the Mechanism for Solution

Additive Effects on the Formation of Bulk Heterojunction Films. Adv. Energy Mater. 2015, 5.

DOI: 10.1002/aenm.201400975.

62. Rogers, J. T.; Schmidt, K.; Toney, M. F.; Bazan, G. C.; Kramer, E. J. Time-Resolved

Structural Evolution of Additive-Processed Bulk Heterojunction Solar Cells. J. Am. Chem. Soc.

2012, 134, 2884–2887.

63. Peet, J.; Kim, J. Y.; Coates, N. E.; Ma, W. L.; Moses, D.; Heeger, A. J.; Bazan, G. C.

Efficiency Enhancement in Low-Bandgap Polymer Solar Cells by Processing with Alkane

Dithiols. Nature Mater. 2007, 6, 497-500.

64. Treat, N. D.; Chabinyc, M. L. Phase Separation in Bulk Heterojunctions of

Semiconducting Polymers and Fullerenes for Photovoltaics. Annu. Rev. Phys. Chem. 2014, 65,

59-81.

65. Grand, C.; Reynolds, J. R. The Interplay Between Structure, Processing, and Properties in

Organic Photovoltaic Devices: How to Translate Recent Laboratory-Scale Developments to

Modules. MRS Communications 2015, 5, 155-167.

Page 53: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

53

66. Lou, S. J.; Szarko, J. M.; Xu, T.; Yu, L.; Marks, T. J.; Chen, L. X. Effects of Additives on

the Morphology of Solution Phase Aggregates Formed by Active Layer Components of High-

Efficiency Organic Solar Cells. J. Am. Chem. Soc. 2011, 133, 20661-20663.

67. Schmidt, K.; Tassone, C. J.; Niskala, J. R.; Yiu, A. T.; Lee, O. P.; Weiss, T. M.; Wang,

C.; Fréchet, J. M. J.; Beaujuge, P. M.; Toney, M. F. A Mechanistic Understanding of Processing

Additive-Induced Efficiency Enhancement in Bulk Heterojunction Organic Solar Cells. Adv.

Mater. 2014, 26, 300-305.

68. Gao, J.; Chen, W.; Dou, L.; Chen, C.-C.; Chang, W.-H.; Liu, Y.; Li, G.; Yang, Y.

Elucidating Double Aggregation Mechanisms in the Morphology Optimization of

Diketopyrrolopyrrole-Based Narrow Bandgap Polymer Solar Cells. Adv. Mater. 2014, 26, 3142-

3147.

69. Yan, H.; Zhu, L.; Li, D.; Zhang, Y.; Yi, Y.; Yang, Y.; Wei, Z.; Brédas, J.-L.

Rationalization of the Selectivity in the Optimization of Processing Conditions for High-

Performance Polymer Solar Cells Based on the Polymer Self-Assembly Ability. J. Phys. Chem.

C 2014, 118, 29473-29481.

70. Shao, M.; Keum, J. K.; Kumar, R.; Chen, J.; Browning, J. F.; Das, S.; Chen, W.; Hou, J.;

Do, C.; Littrell, K. C.; Rondinone, A.; Geohegan, D. B.; Sumpter, B. G.; Xiao, K. Understanding

How Processing Additives Tune the Nanoscale Morphology of High Efficiency Organic

Photovoltaic Blends: From Casting Solution to Spun-Cast Thin Film. Adv. Funct. Mater. 2014,

24, 6647-6657.

Page 54: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

54

71. Brabec, C. J.; Heeney, M.; McCulloch, I.; Nelson, J. Influence of Blend Microstructure

on Bulk Heterojunction Organic Photovoltaic Performance. Chem. Soc. Rev. 2011, 40, 1185-

1199.

72. Wang, H.; Gomez, E. D.; Kim, J.; Guan, Z.; Jaye, C.; Fischer, D. A.; Kahn, A.; Loo, Y.-

L. Device Characteristics of Bulk-Heterojunction Polymer Solar Cells are Independent of

Interfacial Segregation of Active Layers. Chem. Mater. 2011, 23, 2020-2023.

73. Graham, K. R.; Wieruszewski, P. M.; Stalder, R.; Hartel, M. J.; Mei, J.; So, F.; Reynolds,

J. R. Improved Performance of Molecular Bulk-Heterojunction Photovoltaic Cells through

Predictable Selection of Solvent Additives. Adv. Funct. Mater. 2012, 22, 4801-4813.

74. Tsang, S.-W.; Chen, S.; So, F. Energy Level Alignment and Sub-Bandgap Charge

Generation in Polymer:Fullerene Bulk Heterojunction Solar Cells. Adv. Mater. 2013, 25, 2434-

2439.

75. Graham, K. R.; Cabanetos, C.; Jahnke, J. P.; Idso, M. N.; El Labban, A.; Ngongang

Ndjawa, G. O.; Heumueller, T.; Vandewal, K.; Salleo, A.; Chmelka, B. F.; Amassian, A.;

Beaujuge, P. M.; McGehee, M. D. Importance of the Donor:Fullerene Intermolecular

Arrangement for High-Efficiency Organic Photovoltaics. J. Am. Chem. Soc. 2014, 136, 9608-

9618.

76. Chen, S.; Tsang, S.-W.; Lai, T.-H.; Reynolds, J. R.; So, F. Dielectric Effect on the

Photovoltage Loss in Organic Photovoltaic Cells. Adv. Mater. 2014, 26, 6125-6131.

77. Lai, T.-H.; Tsang, S.-W.; Manders, J. R.; Chen, S.; So, F. Properties of Interlayer for

Organic Photovoltaics. Mater. Today 2013, 16, 424-432.

Page 55: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

55

78. Meager, I.; Ashraf, R. S.; Mollinger, S.; Schroeder, B. C.; Bronstein, H.; Beatrup, D.;

Vezie, M. S.; Kirchartz, T.; Salleo, A.; Nelson, J.; McCulloch, I. Photocurrent Enhancement

from Diketopyrrolopyrrole Polymer Solar Cells through Alkyl-Chain Branching Point

Manipulation. J. Am. Chem. Soc. 2013, 135, 11537-11540.

79. Zhang, F.; Hu, Y.; Schuettfort, T.; Di, C.-a.; Gao, X.; McNeill, C. R.; Thomsen, L.;

Mannsfeld, S. C. B.; Yuan, W.; Sirringhaus, H.; Zhu, D. Critical Role of Alkyl Chain Branching

of Organic Semiconductors in Enabling Solution-Processed N-Channel Organic Thin-Film

Transistors with Mobility of up to 3.50 cm2 V–1 s–1. J. Am. Chem. Soc. 2013, 135, 2338-2349.

Page 56: Structure-Property Relationships Directing Transport and ... · 1 Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo Polymers Caroline Grand,1

56

For Table of Contents use only:

Structure-Property Relationships Directing Transport and Charge Separation in Isoindigo

Polymers

Caroline Grand, Sujin Baek, Tzung-Han Lai, Nabankur Deb, Wojciech Zajaczkowski, Romain

Stalder, Klaus Müllen, Wojciech Pisula, David G. Bucknall, Franky So, John R. Reynolds